You are on page 1of 51

Spectra of Atoms and Molecules

May 24, 2005


v
For Robin, Elizabeth, and Victoria
Preface
In this second edition I have mainly updated and revised the material presented in the
rst edition. For example, the 1998 revision of the physical constants has been used
throughout, and the use of symbols and units conforms more closely to recommended
practice. The level of treatment and spirit of the book have not changed. I still aim
to meet the needs of new students of spectroscopy regardless of their background. I
have restrained myself and have not introduced spherical tensors, for example, because
I believe that too many new concepts at one time are confusing.
A certain amount of new material has been added based on my recent experiences
with what is misleadingly called quantitative spectroscopy. Spectroscopists are gen-
erally divided into two camps: those who interpret the spectral positions of lines and
bands, and those who concern themselves more with line and band intensities. The
latter camp is populated mainly by analytical chemists, but includes astronomers and
atmospheric scientists as well.
Nothing in spectroscopy causes as much confusion as line intensities. Some of the
problems seem to originate from the degeneracies inherent in atomic and molecular
systems. The usual intensity formulas are derived (as in Chapter 1) for transitions
between nondegenerate quantum states, while measurements are generally made on
transitions between degenerate energy levels. The correct inclusion of this degeneracy
turns out to be a nontrivial problem and is presented in Chapter 5 for atoms, but the
expressions given there also apply to molecular systems. Even the denition of what
constitutes a line can be a source of diculties.
Line intensities are also confusing because of the dozens of dierent units used to
report line and band strengths. The best procedure is to derive and cite all formulas
in SI units, and then make any needed conversions to customary units in a second
step. It is surprisingly dicult to locate line intensity formulas in SI units, with the
appropriate degeneracies included. The line intensity formulas listed in this book should
prove useful to the modern student.
Other than the addition of material pertaining to line intensities in Chapters 5 to
10, a major change in the second edition is in the discussion of the Raman eect and
light scattering (Chapter 8). The standard theoretical treatment of light scattering and
the Raman eect, as rst presented by Placzek in the 1930s, has been added. Although
Placzeks approach is hardly light reading, the diligent student will nd the derivations
illuminating. A solid understanding of the classical and quantum mechanical theory of
polarizability of molecules is indispensable in the area of nonlinear spectroscopy.
I am very grateful for the comments and helpful criticism from many people, partic-
ularly F. R. McCourt, R. J. Le Roy, C. Bissonette, K. Lehmann, A. Anderson, R. Shiell,
and J. Hardwick. I also thank my fall 2004 graduate class in molecular spectroscopy
(M. Dick, D. Fu, S. Gunal, T. Peng, and S. Yu) for their comments and corrections.
vii
viii
The gures for the second edition have been prepared by S. M. McLeod, T. Nguyen,
Y. Bresler, and E. R. Bernath.
Finally, my wife Robin has made the second edition possible through her continuing
encouragement and understanding. My special thanks to her.
Ontario P.F.B.
August 2004
Preface to First Edition
This book is designed as a textbook to introduce advanced undergraduates and, par-
ticularly, new graduate students to the vast eld of spectroscopy. It presumes that the
student is familiar with the material in an undergraduate course in quantum mechanics.
I have taken great care to review the relevant mathematics and quantum mechanics as
needed throughout the book. Considerable detail is provided on the origin of spectro-
scopic principles. My goal is to demystify spectroscopy by showing the necessary steps
in a derivation, as appropriate in a textbook.
The digital computer has permeated all of science including spectroscopy. The ap-
plication of simple analytical formulas and the nonstatistical graphical treatment of
data are long dead. Modern spectroscopy is based on the matrix approach to quantum
mechanics. Real spectroscopic problems can be solved on the computer more easily if
they are formulated in terms of matrix operations rather than dierential equations.
I have tried to convey the spirit of modern spectroscopy, through the extensive use of
the language of matrices.
The infrared and electronic spectroscopy of polyatomic molecules makes extensive
use of group theory. Rather than assume a previous exposure or try to summarize group
theory in a short chapter, I have chosen to provide a more thorough introduction. My
favorite book on group theory is the text by Bishop, Group Theory and Chemistry, and
I largely follow his approach to the subject.
This book is not a monograph on spectroscopy, but it can be protably read by
physicists, chemists, astronomers, and engineers who need to become acquainted with
the subject. Some topics in this book, such as parity, are not discussed well in any of
the textbooks or monographs that I have encountered. I have tried to take particular
care to address the elementary aspects of spectroscopy that students have found to be
most confusing.
To the uninitiated, the subject of spectroscopy seems enshrouded in layers of bewil-
dering and arbitrary notation. Spectroscopy has a long tradition so many of the symbols
are rooted in history and are not likely to change. Ultimately all notation is arbitrary,
although some notations are more helpful than others. One of the goals of this book is
to introduce the language of spectroscopy to the new student of the subject. Although
the student may not be happy with some aspects of spectroscopic notation, it is easier
to adopt the notation than to try to change long-standing spectroscopic habits.
The principles of spectroscopy are timeless, but spectroscopic techniques are more
transient. Rather than focus on the latest methods of recording spectra (which will be
out of fashion tomorrow), I concentrate on the interpretation of the spectra themselves.
This book attempts to answer the question: What information is encoded in the spectra
of atoms and molecules?
ix
x
A scientic subject cannot be mastered without solving problems. I have therefore
provided many spectroscopic problems at the end of each chapter. These problems have
been acquired over the years from many people including M. Bareld, S. Kukolich, R.
W. Field, and F. McCourt. In addition I have borrowed many problems either directly
or with only small changes from many of the books listed as general references at the
end of each chapter and from the books listed in Appendix D. I thank these people and
apologize for not giving them more credit!
Spectroscopy needs spectra and diagrams to help interpret the spectra. Although the
ultimate analysis of a spectrum may involve the tting of line positions and intensities
with a computer program, there is much qualitative information to be gained by the
inspection of a spectrum. I have therefore provided many spectra and diagrams in this
book. In addition to the specic gure acknowledgments at the end of the appendices, I
would like to thank a very talented group of undergraduates for their eorts. J. Ogilvie,
K. Walker, R. LeBlanc, A. Billyard, and J. Dietrich are responsible for the creation of
most of the gures in this book.
I also would like to thank the many people who read drafts of the entire book or
of various chapters. They include F. McCourt, M. Dulick, D. Klapstein, R. Le Roy, N.
Isenor, D. Irish, M. Morse, C. Jarman, P. Colarusso, R. Bartholomew, and C. Zhao.
Their comments and corrections were very helpful. Please contact me about other errors
in the book and with any comments you would like to make. I thank Heather Hergott
for an outstanding job typing the manuscript.
Finally, I thank my wife Robin for her encouragement and understanding. Without
her this book would never have been written.
Ontario P.F.B.
January 1994
Contents
1 Introduction 3
1.1 Waves, Particles, and Units . . . . . . . . . . . . . . . . . . . . . . . . . 3
1.2 The Electromagnetic Spectrum . . . . . . . . . . . . . . . . . . . . . . . 5
1.3 Interaction of Radiation with Matter . . . . . . . . . . . . . . . . . . . . 7
Blackbody Radiation . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 7
Einstein A and B Coecients . . . . . . . . . . . . . . . . . . . . . . . . 8
Absorption and Emission of Radiation . . . . . . . . . . . . . . . . . . . 11
Beers Law . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 19
Lineshape Functions . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 21
Natural Lifetime Broadening . . . . . . . . . . . . . . . . . . . . . . . . 22
Pressure Broadening . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 27
Doppler Broadening . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 28
Transit-Time Broadening . . . . . . . . . . . . . . . . . . . . . . . . . . 31
Power Broadening . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 33
2 Molecular Symmetry 41
2.1 Symmetry Operations . . . . . . . . . . . . . . . . . . . . . . . . . . . . 41
Operator Algebra . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 42
Symmetry Operator Algebra . . . . . . . . . . . . . . . . . . . . . . . . . 46
2.2 Groups . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 48
Point Groups . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 49
Classes . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 50
Subgroups . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 51
2.3 Notation for Point Groups . . . . . . . . . . . . . . . . . . . . . . . . . . 52
3 Matrix Representation of Groups 58
3.1 Vectors and Matrices . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 58
Matrix Eigenvalue Problem . . . . . . . . . . . . . . . . . . . . . . . . . 63
Similarity Transformations . . . . . . . . . . . . . . . . . . . . . . . . . 65
3.2 Symmetry Operations and Position Vectors . . . . . . . . . . . . . . . . 65
Reection . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 66
Rotation . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 67
Inversion . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 68
Rotation-Reection . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 68
Identity . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 69
3.3 Symmetry Operators and Basis Vectors . . . . . . . . . . . . . . . . . . 69
3.4 Symmetry Operators and Basis Functions . . . . . . . . . . . . . . . . . 72
Function Spaces . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 72
xi
xii CONTENTS
GramSchmidt Procedure . . . . . . . . . . . . . . . . . . . . . . . . . . 74
Transformation Operators . . . . . . . . . . . . . . . . . . . . . . . . . . 74
3.5 Equivalent, Reducible, and Irreducible Representations . . . . . . . . . . 77
Equivalent Representations . . . . . . . . . . . . . . . . . . . . . . . . . 77
Unitary Representations . . . . . . . . . . . . . . . . . . . . . . . . . . . 78
Reducible and Irreducible Representations . . . . . . . . . . . . . . . . . 78
3.6 Great Orthogonality Theorem . . . . . . . . . . . . . . . . . . . . . . . . 79
Characters . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 81
3.7 Character Tables . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 83
Mulliken Notation . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 84
4 Quantum Mechanics and Group Theory 91
4.1 Matrix Representation of the Schrodinger Equation . . . . . . . . . . . . 91
4.2 BornOppenheimer Approximation . . . . . . . . . . . . . . . . . . . . . 97
4.3 Symmetry of the Hamiltonian Operator . . . . . . . . . . . . . . . . . . 99
4.4 Projection Operators . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 100
4.5 Direct Product Representations . . . . . . . . . . . . . . . . . . . . . . . 103
4.6 Integrals and Selection Rules . . . . . . . . . . . . . . . . . . . . . . . . 105
5 Atomic Spectroscopy 109
5.1 Background . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 109
5.2 Angular Momentum . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 111
5.3 The Hydrogen Atom and One-Electron Spectra . . . . . . . . . . . . . . 115
Vector Model . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 118
Spin-Orbit Coupling . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 120
5.4 Many-Electron Atoms . . . . . . . . . . . . . . . . . . . . . . . . . . . . 124
5.5 Selection Rules . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 134
5.6 Atomic Spectra . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 137
Hyperne Structure . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 137
Hydrogen Atom . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 138
5.7 Intensity of Atomic Lines . . . . . . . . . . . . . . . . . . . . . . . . . . 140
5.8 Zeeman Eect . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 147
PaschenBack Eect . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 152
5.9 Stark Eect . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 153
6 Rotational Spectroscopy 161
6.1 Rotation of Rigid Bodies . . . . . . . . . . . . . . . . . . . . . . . . . . . 161
6.2 Diatomic and Linear Molecules . . . . . . . . . . . . . . . . . . . . . . . 169
Selection Rules . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 171
Centrifugal Distortion . . . . . . . . . . . . . . . . . . . . . . . . . . . . 174
Vibrational Angular Momentum . . . . . . . . . . . . . . . . . . . . . . . 177
6.3 Line Intensities for Diatomic and Linear Molecules . . . . . . . . . . . . 182
6.4 Symmetric Tops . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 185
Molecule and Space-Fixed Angular Momenta . . . . . . . . . . . . . . . 185
Rotational Spectra . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 190
Centrifugal Distortion . . . . . . . . . . . . . . . . . . . . . . . . . . . . 190
Line Intensity . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 192
6.5 Asymmetric Tops . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 193
Selection Rules . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 195
Line Intensity . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 198
CONTENTS xiii
6.6 Structure Determination . . . . . . . . . . . . . . . . . . . . . . . . . . . 198
7 Vibrational Spectroscopy 208
7.1 Diatomic Molecules . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 208
Wavefunctions for Harmonic and Anharmonic Oscillators . . . . . . . . 215
Vibrational Selection Rules for Diatomics . . . . . . . . . . . . . . . . . 216
Dissociation Energies from Spectroscopic Data . . . . . . . . . . . . . . 220
Vibration-Rotation Transitions of Diatomics . . . . . . . . . . . . . . . . 223
Combination Dierences . . . . . . . . . . . . . . . . . . . . . . . . . . . 224
7.2 Vibrational Motion of Polyatomic Molecules . . . . . . . . . . . . . . . . 226
Classical Mechanical Description . . . . . . . . . . . . . . . . . . . . . . 226
Quantum Mechanical Description . . . . . . . . . . . . . . . . . . . . . . 231
Internal Coordinates . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 232
Symmetry Coordinates . . . . . . . . . . . . . . . . . . . . . . . . . . . . 233
Symmetry of Normal Modes . . . . . . . . . . . . . . . . . . . . . . . . . 238
Selection Rules for Vibrational Transitions . . . . . . . . . . . . . . . . 245
Vibration-Rotation Transitions of Linear Molecules . . . . . . . . . . . . 247
Nuclear Spin Statistics . . . . . . . . . . . . . . . . . . . . . . . . . . . . 253
Excited Vibrational States of Linear Molecules . . . . . . . . . . . . . . 256
7.3 Vibrational Spectra of Symmetric Tops . . . . . . . . . . . . . . . . . . 259
Coriolis Interactions in Molecules . . . . . . . . . . . . . . . . . . . . . . 260
7.4 Infrared Transitions of Spherical Tops . . . . . . . . . . . . . . . . . . . 266
7.5 Vibrational Spectra of Asymmetric Tops . . . . . . . . . . . . . . . . . . 270
7.6 Vibration-Rotation Line Intensities . . . . . . . . . . . . . . . . . . . . . 272
Line Intensity Calculations . . . . . . . . . . . . . . . . . . . . . . . . . 275
7.7 Fermi and Coriolis Perturbations . . . . . . . . . . . . . . . . . . . . . . 278
7.8 Inversion Doubling and Fluxional Behavior . . . . . . . . . . . . . . . . 280
8 Light Scattering and the Raman Eect 293
8.1 Background . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 293
Classical Model . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 293
Quantum Model . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 298
Polarization . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 305
8.2 Rotational Raman Eect . . . . . . . . . . . . . . . . . . . . . . . . . . . 307
Diatomic Molecules . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 309
8.3 Vibration-Rotation Raman Spectroscopy . . . . . . . . . . . . . . . . . . 309
Diatomic Molecules . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 309
8.4 Rayleigh and Raman Intensities . . . . . . . . . . . . . . . . . . . . . . . 310
Classical Theory . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 310
Vibrational Intensity Calculations . . . . . . . . . . . . . . . . . . . . . 315
8.5 Conclusions . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 317
9 Electronic Spectroscopy of Diatomics 321
9.1 Orbitals and States . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 321
9.2 Vibrational Structure . . . . . . . . . . . . . . . . . . . . . . . . . . . . 326
9.3 Rotational Structure of Diatomic Molecules . . . . . . . . . . . . . . . . 332
Singlet-Singlet Transitions . . . . . . . . . . . . . . . . . . . . . . . . . . 332
Nonsinglet Transitions . . . . . . . . . . . . . . . . . . . . . . . . . . . . 337
9.4 The Symmetry of Diatomic Energy Levels: Parity . . . . . . . . . . . . . 346
Total (+/) Parity . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 347
xiv CONTENTS
Rotationless (e/f ) Parity . . . . . . . . . . . . . . . . . . . . . . . . . . 349
Gerade/Ungerade (g/u) Parity . . . . . . . . . . . . . . . . . . . . . . . 351
Symmetric/Antisymmetric (s/a) Parity . . . . . . . . . . . . . . . . . . 352
9.5 Rotational Line Intensities . . . . . . . . . . . . . . . . . . . . . . . . . . 353
9.6 Dissociation, Photodissociation, and Predissociation . . . . . . . . . . . 359
10 Electronic Spectroscopy of Polyatomics 367
10.1 Orbitals and States . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 367
Walshs Rules: Qualitative Molecular Orbital Theory . . . . . . . . . . . 368
H uckel Molecular Orbital Theory . . . . . . . . . . . . . . . . . . . . . . 372
10.2 Vibrational Structure of Electronic Transitions . . . . . . . . . . . . . . 379
10.3 Vibronic Coupling: The HerzbergTeller Eect . . . . . . . . . . . . . . 382
10.4 JahnTeller Eect . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 385
10.5 RennerTeller Eect . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 386
10.6 Nonradiative Transitions: Jablonski Diagram . . . . . . . . . . . . . . . 388
10.7 Photoelectron Spectroscopy . . . . . . . . . . . . . . . . . . . . . . . . . 390
10.8 Rotational Structure: H
2
CO and HCN . . . . . . . . . . . . . . . . . . . 391
10.9 Intensity of Transitions . . . . . . . . . . . . . . . . . . . . . . . . . . . . 398
A Units, Conversions, and Physical Constants 406
B Character Tables 408
C Direct Product Tables 420
D Introductory Textbooks 424
Figure Acknowledgments 427
Index 431
Spectra of Atoms and Molecules
Chapter 1
Introduction
1.1 Waves, Particles, and Units
Spectroscopy is the study of the interaction of light with matter. To begin, a few words
about light, matter, and the eect of light on matter are in order.
Light is an electromagnetic wave represented (for the purposes of this book) by the
plane waves
E(r, t) = E
0
cos(k r t +
0
) (1.1)
or
E(r, t) = Re
_
E
0
e
i(krt+
0
)
_
. (1.2)
In this book vectors and matrices are written in bold Roman type, except in certain
gures in which vectors are indicated with a half arrow (e.g.,

E) for clarity. There
is an electric eld E (in volts per meter) perpendicular to k that propagates in the
direction k and has an angular frequency = 2 = 2/T. The frequency (in hertz)
is the reciprocal of the period T (in seconds), that is, = 1/T. The period T and the
wavelength are dened in Figure 1.1 with k in the z direction. The wavevector k
has a magnitude |k| = k = 2/ and a direction given by the normal to the plane of
constant phase. |E
0
| is the amplitude of the electric eld, while kr t +
0
is the
phase (
0
is an initial phase angle of arbitrary value).
The presence of a magnetic eld, also oscillating at angular frequency and orthog-
onal to both E and k, is ignored in this book. Other complications such as Maxwells
equations, Gaussian laser beams, birefringence, and vector potentials are also not con-
sidered. These subjects, although part of spectroscopy in general, are discussed in books
on optics, quantum optics, lasers, or electricity and magnetism.
Wavelength and frequency are related by the equation = c, in which c is the
speed of the electromagnetic wave. In vacuum c = c
0
, but in general c = c
0
/n with n as
the index of refraction of the propagation medium. Since has the same value in any
medium, the wavelength also depends on the index of refraction of the medium. Thus
since
= c (1.3)
we must have
3
4 1. Introduction
t
T
E
E0
z
E
E0
l
Figure 1.1: The electric eld at t = 0 as a function of z is plotted in the upper panel, while
the lower panel is the corresponding plot at z = 0 as a function of time.

0
n
=
c
0
n
. (1.4)
Historically, direct frequency measurements were not possible in the infrared and
visible regions of the spectrum. It was therefore convenient to measure (and report)
in air, correct for the refractive index of air to give
0
, and then dene = 1/
0
, with

0
in cm. Before SI units were adopted, the centimeter was more widely used than the
meter so that represents the number of wavelengths in one centimeter in vacuum and
as a consequence is called the wavenumber. The units of the wavenumber are cm
1
(reciprocal centimeters) but common usage also calls the cm
1
the wavenumber.
Fortunately, the SI unit for , the m
1
, is almost never used so that this sloppy, but
standard, practice causes no confusion.
The oscillating electric eld is a function of both spatial (r) and temporal (t) vari-
ables. If the direction of propagation of the electromagnetic wave is along the z-axis
and the wave is examined at one instant of time, say t = 0, then for
0
= 0,
E = E
0
cos(kz) = E
0
cos
2z

. (1.5)
Alternatively, the wave can be observed at a single point, say z = 0, as a function of
time
E = E
0
cos(t) = E
0
cos(2t). (1.6)
Both equations (1.5) and (1.6) are plotted in Figure 1.1 with arbitrary initial phases.
1.2 The Electromagnetic Spectrum 5
In contrast to longitudinal waves, such as sound waves, electromagnetic waves are
transverse waves. If the wave propagates in the z direction, then there are two possible
independent transverse directions, x and y. This leads to the polarization of light, since
E could lie either along x or along y, or more generally, it could lie anywhere in the
xy-plane. Therefore we may write
E = E
x

i + E
y

j, (1.7)
with

i and

j representing unit vectors lying along the x- and y-axes.


The wave nature of light became rmly established in the nineteenth century, but
by the beginning of the twentieth century, light was also found to have a particle aspect.
The wave-particle duality of electromagnetic radiation is dicult to visualize since there
are no classical, macroscopic analogs. In the microscopic world, electromagnetic waves
seem to guide photons (particles) of a denite energy E and momentum p with
E = h = =
hc

= 10
2
hc (1.8)
and
p =
h

= k. (1.9)
The factor of 10
2
in equation (1.8) comes from the conversion of cm
1
for into m
1
.
In 1924 it occurred to de Broglie that if electromagnetic waves could display prop-
erties associated with particles, then perhaps particles could also display wavelike prop-
erties. Using equation (1.9), he postulated that a particle should have a wavelength,
=
h
p
=
h
mv
. (1.10)
This prediction of de Broglie was veried in 1927 by Davisson and Germers observation
of an electron beam diracted by a nickel crystal.
In this book SI units and expressions are used as much as possible, with the tra-
ditional spectroscopic exceptions of the angstrom (

A) and the wavenumber (cm


1
).
The symbols and units used will largely follow the International Union of Pure and
Applied Chemistry (IUPAC) recommendations of the Green Book by I. M. Mills et
al.
1
The fundamental physical constants, as supplied in Appendix A, are the 1998 Mohr
and Taylor
2
values. Notice that the speed of light in vacuum (c
0
) is xed exactly at
299 792 458 m/s. The atomic masses used are the 2003 Audi, Wapstra, and Thibault
3
values and atomic mass units have the recommended symbol, u.
1
1.2 The Electromagnetic Spectrum
There are traditional names associated with the various regions of the electromagnetic
spectrum. The radio frequency region (3 MHz3 GHz) has photons of sucient energy
to ip nuclear spins (nuclear magnetic resonance (NMR)) in magnetic elds of a few
tesla. In the microwave region (3 GHz3 000 GHz) energies correspond to rotational
transitions in molecules and to electron spin ips (electron spin resonance (ESR)). Un-
like all the spectra discussed in this book, NMR and ESR transitions are induced by
6 1. Introduction
Figure 1.2: The electromagnetic spectrum.
the oscillating magnetic eld of the electromagnetic radiation. Infrared quanta (100
cm
1
13 000 cm
1
) excite the vibrational motion in matter. Visible and ultraviolet
(UV) transitions (10 000

A100

A) involve valence electron rearrangements in mole-
cules (1 nm = 10

A). Core electronic transitions are promoted at x-ray wavelengths
(100

A0.1

A). Finally, below 0.1

A in wavelength, -rays are associated with nuclear
processes. Chemists customarily use the units of MHz or GHz for radio and microwave
radiation, cm
1
for infrared radiation, and nm or

A for visible, UV, and x-ray ra-
diation (Figure 1.2). These customary units are units of frequency (MHz), reciprocal
wavelength (cm
1
), and wavelength (

A).
It is worth noting that the dierent regions of the spectrum do not possess sharp
borders and that the type of molecular motion associated with spectroscopy in each
region is only approximate. For example, overtone vibrational absorption can be found
in the visible region of the spectrum (causing the blue color of the oceans). Infrared
electronic transitions are also not rare, for example, the BallikRamsay electronic tran-
sition of C
2
.
A further subdivision of the infrared, visible, and ultraviolet regions of the spectrum
is customary. The infrared region is divided into the far-infrared (33333 cm
1
), mid-
infrared (3333 333 cm
1
), and near-infrared (3 33313 000 cm
1
) regions. In the far-
infrared region are found rotational transitions of light molecules, phonons of solids,
and metal-ligand vibrations, as well as ring-puckering and torsional motions of many
organic molecules. The mid-infrared is the traditional infrared region in which the
fundamental vibrations of most molecules lie. The near-infrared region is associated
with overtone vibrations and a few electronic transitions. The visible region is divided
into the colors of the rainbow from the red limit at about 7 800

A to the violet at 4 000

A. The near-ultraviolet region covers 4 000



A2 000

A, while the vacuum ultraviolet
region is 2 000

A100

A. The vacuum ultraviolet region is so named because air is
opaque to wavelengths below 2 000

A, so that only evacuated instruments can be used
when spectra are taken in this region.
1.3 Interaction of Radiation with Matter 7
It is a spectroscopic custom to report all infrared, visible, and near-ultraviolet
wavelengths as air wavelengths (), rather than as vacuum wavelengths (
0
). Of course,
below 2000

A all wavelengths are vacuum wavelengths since measurements in air are
not possible. The wavenumber is related to energy, E = 10
2
hc , and is the reciprocal
of the vacuum wavelength in centimeters, = 1/
0
, but in air = 1/
0
= 1/n.
For accurate work, it is necessary to correct for the refractive index of air. This can
be seen, for example, by considering dry air at 15

C and 760 Torr for which n =


1.000 278 1 at 5 000

A.
4
Thus = 5 000.000

A in air corresponds to
0
= 5 001.391

A
in vacuum and = 19 994.44 cm
1
rather than 20 000 cm
1
!
1.3 Interaction of Radiation with Matter
Blackbody Radiation
The spectrum of the radiation emitted by a blackbody is important both for historical
reasons and for practical applications. Consider a cavity (Figure 1.3) in a material that
is maintained at constant temperature T. The emission of radiation from the cavity
walls is in equilibrium with the radiation that is absorbed by the walls. It is convenient
to dene a radiation density (with units of joules/m
3
) inside the cavity. The frequency
distribution of this radiation is represented by the function

, which is the radiation


density in the frequency interval between and + d (Figure 1.4), and is dened so
that
=
_

0

d. (1.11)
Therefore, the energy density function

has units of joule-seconds per cubic meter (J


s m
3
). The distribution function characterizing the intensity of the radiation emitted
from the hole is labeled I

(units of watt-seconds per square meter of the hole). In the


radiometric literature the quantity I =
_
I

d (W m
2
) would be called the radiant
excitance and I

(W s m
2
) would be the spectral radiant excitance.
5
The recommended
radiometric symbol for excitance is M, which is not used here because of possible
confusion with the symbol for dipole moment. The functions

and I

are universal
functions depending only upon the temperature and frequency, and are independent of
the shape or size of the cavity and of the material of construction as long as the hole
is small.
Planck obtained the universal function,

(T) =
8h
3
c
3
1
e
h/kT
1
, (1.12)
named in his honor. The symbol k = 1.380 650 310
23
J K
1
(Appendix A) in equation
(1.12) represents the Boltzmann constant. Geometrical considerations (Problem 13)
then give the relationship between I

and

as
I

c
4
. (1.13)
Figure 1.5. shows

as a function of and the dependence on the temperature T.


8 1. Introduction
Figure 1.3: Cross section of a blackbody cavity at a temperature T with a radiation density

emitting radiation with intensity I

from a small hole.


Figure 1.4: The Planck function

() is a distribution function dened by d/d =

() or
=

d.
Einstein A and B Coecients
Consider a collection of N two-level systems (Figure 1.6) in a volume of 1 m
3
with
upper energy E
1
and lower energy E
0
, all at a constant temperature T and bathed by
the radiation density

(T). Since the entire collection is in thermal equilibrium, if the


number of systems with energy E
1
is N
1
and the number of systems with energy E
0
is
N
0
, then the populations N
1
and N
0
(N = N
1
+ N
0
) are necessarily related by
1.3 Interaction of Radiation with Matter 9
Figure 1.5: The Planck function at 77 K, 200 K, and 300 K.
Figure 1.6: A two-level system.
N
1
N
0
= e
h
10
/kT
, (1.14)
in which h
10
= E
1
E
0
. This is the well-known Boltzmann expression for thermal
equilibrium between nondegenerate levels.
10 1. Introduction
Figure 1.7: Schematic representations of absorption (top), spontaneous emission (middle) and
stimulated emission (bottom) processes in a two-level system.
There are three possible processes that can change the state of the system from E
0
to E
1
or from E
1
to E
0
: absorption, spontaneous emission, and stimulated emission
(Figure 1.7). Absorption results from the presence of a radiation density

(
10
) of the
precise frequency needed to drive a transition from the ground state to the excited state
at the rate
dN
1
dt
= B
10

(
10
)N
0
. (1.15)
The coecient B
10
is thus a rate constant and is known as the Einstein absorption
coecient or Einstein B coecient. Similarly if the system is already in an excited
state, then a photon of energy h
10
(provided by

) can induce the system to make


the transition to the ground state. The rate for stimulated emission is given by
dN
1
dt
= B
10

(
10
)N
1
, (1.16)
in which B
10
is the stimulated emission coecient. Finally the system in the excited
state can spontaneously emit a photon at a rate
dN
1
dt
= A
10
N
1
. (1.17)
Since the system is at equilibrium, the rate of population of the excited state by
absorption must balance the rate of depopulation by stimulated and spontaneous emis-
sion, so that
1.3 Interaction of Radiation with Matter 11
N
0
B
10

= A
10
N
1
+ B
10

N
1
(1.18)
and hence using (1.14)
N
1
N
0
=
B
10

A
10
+ B
10

= e
h
10
/kT
. (1.19)
Solving for

in equation (1.19) then yields

(
10
) =
A
10
B
10
e
h
10
/kT
B
10
. (1.20)
However,

(
10
) is also given by the Planck function (1.12)

(
10
) =
8h
3
10
c
3
1
e
h
10
/kT
1
.
For expressions (1.12) and (1.20) both to be valid, it is necessary that
B
10
= B
10
(1.21)
and that
A
10
=
8h
3
10
c
3
B
10
. (1.22)
Remarkably, the rate constants for absorption and stimulated emissiontwo apparently
dierent physical processesare identical (1.21). Moreover, the spontaneous emission
rate (lifetime) can be determined from the absorption coecient (1.22). Note, however,
the
3
10
factor in (1.22), which plays an important role in the competition between the
induced and spontaneous emission processes.
Absorption and Emission of Radiation
The interaction of electromagnetic radiation with matter can be described by a simple
semiclassical model. In the semiclassical treatment the energy levels of molecules are
obtained by solution of the time-independent Schrodinger equation

H
n
= E
n

n
,
while the electromagnetic radiation is treated classically. Consider a two-level system
described by lower and upper state wavefunctions,
0
and
1
(Figure 1.8), respectively.
Electromagnetic radiation that fullls the Bohr condition, E
1
E
0
= h = , is
applied to the system in order to induce a transition from the lower energy state at
E
0
to the upper energy state at E
1
. The molecule consists of nuclei and electrons at
positions r
i
possessing charges q
i
. The system as a whole thus has a net dipole moment
with Cartesian components

x
=

x
i
q
i
, (1.23)

y
=

y
i
q
i
, (1.24)
12 1. Introduction
Figure 1.8: Two-level system.
Figure 1.9: Three particles with charges q
1
, q
2
, and q
3
located at positions r
1
, r
2
, and r
3
.

z
=

z
i
q
i
, (1.25)
where (x
i
, y
i
, z
i
) give the coordinates of particle i relative to the center of mass of the
molecule (Figure 1.9).
The interaction of the radiation with the material system is taken into account by
the addition of the time-dependent perturbation (see Chapter 4 and the rst section of
this chapter for denitions),

(t) = E(t)
= E
0
cos (k r t). (1.26)
If the oscillating electric eld is in the z direction (E
0
= (0, 0, E
0z
)) and the system is
at the origin, r = 0 (the wavelength is greater than the dimensions of the system to
avoid having dierent electric eld strengths at dierent parts of the molecule), then

(t) =
z
E
0z
cos (t) (1.27)
= E cos (t). (1.28)
The transition probability is obtained by solving the time-dependent Schrodinger equa-
tion
1.3 Interaction of Radiation with Matter 13
i

t
= (

H +

H

(t)). (1.29)
In the absence of

H

, the two time-dependent solutions of equation (1.29) are

0
(t) =
0
e
iE
0
t/
=
0
e
i
0
t

1
(t) =
1
e
iE
1
t/
=
1
e
i
1
t
,
with
i
= E
i
/. An uppercase Greek is used for a time-dependent wavefunction,
while a lowercase represents a time-independent wavefunction.
The wavefunction for the perturbed two-level system is given by the linear combi-
nation of the complete set of functions
0
and
1
:
(t) = a
0
(t)
0
e
iE
0
t/
+ a
1
(t)
1
e
iE
1
t/
= a
0

0
e
i
0
t
+ a
1

1
e
i
1
t
, (1.30)
where a
0
and a
1
are time-dependent coecients. Substitution of the solution (1.30) into
the time-dependent Schrodinger equation (1.29) leads to the equation
i( a
0

0
e
i
0
t
+ a
1

1
e
i
1
t
) =

H

a
0

0
e
i
0
t
+

H

a
1

1
e
i
1
t
, (1.31)
where the dot notation a
0
= da
0
/dt is used to indicate derivatives with respect to time.
Multiplication by

0
e
i
0
t
, or

1
e
i
1
t
, followed by integration over all space then
gives two coupled dierential equations
i a
0
= a
0

0
|

|
0
+ a
1

0
|

|
1
e
i
10
t
(1.32a)
i a
1
= a
0

1
|

|
0
e
i
10
t
+ a
1

1
|

|
1
(1.32b)
using the Dirac bracket notation f
1
|

A|f
3
=
_
f

1

Af
3
d. At this stage no approxima-
tions have been made (other than the restriction to the two states
1
and
0
), and
the two equations (1.32) are entirely equivalent to the original Schrodinger equation.
Now if

H

is taken as E cos (t) in the electric-dipole approximation, then the time-


dependent perturbation

H

has odd parity (i.e., is an odd function of the spatial coordi-


nates, see also Chapters 5 and 9). In other words

H

is an odd function since = ez,


while the products |
1
|
2
or |
0
|
2
are even functions so that the integrands

1

H

1
and

0

H

0
are also odd functions. All atomic and molecular states that have denite
parities (either even or odd) with respect to inversion in the space-xed coordinate
system (see Chapters 5 and 9 for further details) have
0
|

|
0
=
1
|

|
1
= 0,
and equations (1.32) reduce to
i a
0
= a
1
M
01
Ee
i
10
t
cos t, (1.33a)
i a
1
= a
0
M
01
Ee
i
10
t
cos t. (1.33b)
The integral M
01
= M
10
=
1
||
0
is the transition dipole moment and is the most
critical factor in determining selection rules and line intensities. In general M
10
is a
vector quantity and the symbol
10
( M
10
) is often used. It is convenient to dene
14 1. Introduction

R
=
M
10
E

, (1.34)
which is known as the Rabi frequency, and to use the identity
cos (t) =
e
it
+ e
it
2
to rewrite equations (1.33a) and (1.33b) as
a
0
=
ia
1

R
(e
i(
10
)t
+ e
i(
10
+)t
)
2
(1.35a)
a
1
=
ia
0

R
(e
i(
10
)t
+ e
i(
10
+)t
)
2
. (1.35b)
The physical meaning of the Rabi frequency will become clear later in this section.
At this stage an approximation can be made by noting that
10
since the system
with Bohr angular frequency (E
1
E
0
)/ =
10
is resonant or nearly resonant with the
optical angular frequency = 2. The terms e
i(
10
)t
and e
i(
10
)t
thus represent
slowly varying functions of time compared to the rapidly oscillating nonresonant terms
e
i(
10
+)t
and e
i(
10
+)t
.
In what is known as the rotating wave approximation the nonresonant high-
frequency terms can be neglected as their eects essentially average to zero because
they are rapidly oscillating functions of time. Upon dening as
10
, equations
(1.35a) and (1.35b) become
a
0
=
i
R
e
it
a
1
2
(1.36a)
a
1
=
i
R
e
it
a
0
2
. (1.36b)
Equations (1.36a) and (1.36b) can be solved analytically. The dierence is often
referred to as the detuning frequency, since it measures how far the electromagnetic
radiation of angular frequency is tuned away from the resonance frequency
10
. The
solution (see Problem 14) to these two simultaneous rst-order dierential equations
with initial conditions a
0
(0) = 1 and a
1
(0) = 0 for the system initially in the ground
state at t = 0 is
a
0
(t) =
_
cos
_
t
2
_
i
_

_
sin
_
t
2
__
e
it/2
(1.37)
and
a
1
(t) = i
_

_
sin
_
t
2
_
e
it/2
, (1.38)
in which = ((
R
)
2
+
2
)
1/2
. These solutions can be checked by substitution into
equations (1.36).
The time-dependent probability that the system will be found in the excited state
is given by
|a
1
(t)|
2
=

2
R

2
sin
2
_
t
2
_
, (1.39)
1.3 Interaction of Radiation with Matter 15
Figure 1.10: The probability for nding the driven two-level system in the excited state for
three detunings: = 0, =
R
, and = 3
R
.
while the corresponding time-dependent probability that the system will be found in
the ground state is given by
|a
0
|
2
= 1 |a
1
|
2
= 1

2
R

2
sin
2
_
t
2
_
. (1.40)
At resonance = 0 and =
R
, so that in this case
|a
1
|
2
= sin
2
_

R
t
2
_
(1.41)
|a
0
|
2
= 1 sin
2
_

R
t
2
_
= cos
2
_

R
t
2
_
. (1.42)
The transition probability |a
1
|
2
is plotted in Figure 1.10 for three detuning frequencies.
The meaning of the Rabi frequency becomes clear from equations (1.34), (1.41),
and Figure 1.10. The system is coherently cycled (i.e., with no abrupt changes in the
phases or amplitudes of the wavefunctions) between the ground and excited state by the
electromagnetic radiation. At resonance the system is completely inverted after a time
t

= /
R
, while o-resonance there is a reduced probability for nding the system in
the excited state.
This simple picture of a coherently driven system has ignored all decay processes
such as spontaneous emission from the excited state. Spontaneous emission of a photon
would break the coherence of the excitation and reset the system to the ground state
(this is referred to as a T
1
process). Similarly, collisions can also cause relaxation in the
system. In fact, collisions can reset the phase of the atomic or molecular wavefunction
(only the relative phases of
1
and
0
are important) without changing any of the
populations (this is referred to as a T
2
process). These phase-changing collisions also
interrupt the coherent cycling of the system. Such processes were rst studied in NMR
(which is the source of the names T
1
and T
2
processes) and are now extensively studied
in the eld of quantum optics.
16 1. Introduction
Figure 1.11: The driven two-level system saturates when relaxation processes (damping) are
included.
The eect of collisions and other relaxation phenomena is to damp out the coherent
cycling of the excited system (called Rabi oscillations). However, Rabi oscillations can
be observed in any quantum system simply by increasing the intensity of the radiation.
This increases the applied electric eld E so that at some point the Rabi cycling fre-
quency exceeds the relaxation frequency,
R

relaxation
, and coherent behavior will
be observed. This is easily achieved in NMR where spin relaxation processes are slow
and many watts of radio frequency power can be applied to the system. In the infrared
and visible region of the spectrum relaxation processes are much faster and Rabi oscil-
lations are normally damped. For example, a real system would oscillate briey when a
strong eld is applied suddenly to it, but it soon loses coherence and saturates (Figure
1.11). When the system is saturated, half of the molecules in the system are in the lower
state and half are in the upper state. The rate of stimulated emission (down) matches
the rate of absorption (up).
For example, consider a 1-W laser beam, 1 mm in diameter interacting with a two-
level system that has a transition dipole moment of 1 debye (1 D = 3.335 64 10
30
C m). What is the Rabi frequency? The intensity of the laser beam is 1.3 10
6
W/m
2
and the electric eld E = |E
0
| is calculated from
I =
_
1
2

0
E
2
_
c (1.43)
with
0
= 8.854 187 810
12
C
2
N
1
m
2
(Appendix A) being the permittivity of free
space, from which the electric eld can be obtained as
E = 27.4

I = 3.1 10
4
V/m
and

R
=
E

= 9.8 10
8
rad/s or
R
= 156 MHz,
1

R
= 6.4 ns.
1.3 Interaction of Radiation with Matter 17
Recall that
0
appears in Coulombs law for the magnitude of the force of interaction
between two electrical charges,
F =
q
1
q
2
4
0
r
2
.
The factor 1/2 in equation (1.43) applies when the radiation is polarized and the elec-
tric eld is given by equation (1.1). Since a typical electronic transition may have a
natural lifetime of 10 ns (
natural
= 6.3 10
8
rad/s), the eects of Rabi cycling are
already present at 1 W. At the megawatt or higher power levels of typical pulsed lasers,
the coherent eects of strong radiation are even more pronounced, provided electri-
cal breakdown is avoided. At these high electric eld strengths, however, the simple
two-level model is not a good description of an atomic or molecular system.
There is much confusion between the various terms and symbols used in the subeld
of radiometry. For example, in physics the commonly used term intensity, I (W m
2
)
of a laser beam would be called the irradiance in radiometry.
5
In radiometry the
terms intensity, I (W sr
1
), and spectral intensity I

= dI/d (W s sr
1
) are instead
used for power per steradian and power per steradian per hertz (respectively). (A
sphere subtends a solid angle of 4 steradians.) In radiometry a distinction is also
made between the excitance, M (W m
2
), of power leaving a surface (e.g., equation
(1.13)) and the irradiance, E (W m
2
), of power falling on or crossing a surface (e.g.,
equation (1.43)) although they have the same dimensions. The right subscript is used
to distinguish between integrated quantities such as the irradiance, E =
_
E

d and
the spectral irradiance, E

= dE/d. In this book we follow the physics custom of using


the single term intensity, I (or I

), for the excitance (or spectral excitance) and the


irradiance (or spectral irradiance), with I =
_
I

d. The term radiance, L =


_
L

d, is
universally reserved for power per square meter per steradian (W m
2
sr
1
), and the
spectral radiance, L

, has dimensions W s m
2
sr
1
in SI units. The spectral radiance
of a blackbody is given by
L
BB

c
4
=
2h
3
c
2
1
e
h/kT
1
. (1.44)
The case of weak electromagnetic radiation interacting with the system is also com-
mon. In fact before the development of the laser in 1960 the weak-eld case applied to all
regions of the spectrum except the radio frequency and microwave regions, for which
powerful coherent sources were available. In the weak-eld case there is a negligible
buildup of population in the excited state, so that a
1
0, a
0
1, and
a
1
=
i
R
2
e
it
. (1.45)
Equation (1.45) is readily integrated to give
a
1
=
i
R
2
_
t
0
e
it
dt
=

R
2
(e
it
1). (1.46)
The probability for nding the system in the excited state after a time t is then obtained
from equation (1.46) as
18 1. Introduction
P
10
= |a
1
|
2
=

2
R

2
sin
2
_
t
2
_
=

2
10
E
2

2
sin
2
((
10
)t/2)
(
10
)
2
. (1.47)
This formula is very deceptive because it assumes monochromatic radiation and short
interaction times. These two requirements are inconsistent with one other because the
Heisenberg energy-time uncertainty principle
Et or t
1
2
(1.48)
must always be satised. If monochromatic radiation is applied to the system for a time
t, then the system sees radiation of width = 1/(2t) in frequency space (this
is certainly not monochromatic!). For example, a pulse of radiation 10 ns long has an
intrinsic width of at least 160 MHz in frequency space.
Before equation (1.47) can be used, the eects of the nite frequency spread of
the radiation must be included. Consider the radiation applied to the system to be
broadband rather than monochromatic and to have a radiation density =
0
E
2
/2.
The total transition probability is given by integrating over all frequencies, that is, by
P
10
=
2
2
10

2
_

()
sin
2
((
10
)t/2)
(
10
)
2
d
=
2
2
10

(
10
)
_
sin
2
((
10
)t/2)
(
10
)
2
d
=

2
10

(
10
)t, (1.49)
in which () is assumed to be slowly varying near
10
so that it can be removed from
the integration. This is indeed the case as sin
2
((
10
)t/2) /(
10
)
2
is sharply
peaked at =
10
(see Figure 1.26). The absorption rate per molecule is thus given by
dP
10
dt
=

2
10

(
10
). (1.50)
In order to derive an expression for the absorption coecient in terms of the tran-
sition dipole moment, equation (1.50) needs to be compared with equation (1.15), with
N
0
N for the weak-eld case: dividing by N then gives the transition probability per
molecule as
d(N
1
/N)
dt
= B
10

(
10
),
or as
dP
10
dt
= B
10

(
10
). (1.51)
A factor 3 is missing from equation (1.50) because equation (1.51) has been derived
using isotropic radiation traveling in x, y, and z directions, while equation (1.50) has
been derived using a plane wave traveling in the z direction. Since only the z component
of the isotropic radiation is eective in inducing a transition, and since () = 2(),
we have
1.3 Interaction of Radiation with Matter 19
B
10
=
1
6
0

2
10
=
2
2
3
0
h
2

2
10
(1.52)
and
A
10
=
16
3

3
3
0
hc
3

2
10
. (1.53)
These equations, (1.52) and (1.53), are key results because they relate the observed
macroscopic transition rates to the microscopic transition dipole moment of an atom or
molecule. Upon substitution of the values of the constants, we obtain A
10
= 3.136
10
7
( )
3

2
10
, with expressed in cm
1
and
10
in debye. Although these equations are
essentially correct, one factor that has been ignored is the possibility of relaxation.
Collisions or the spontaneous radiative lifetime of the upper state have all been
ignored so far. When these losses are considered, the molecular absorption lineshape
changes from a Dirac delta function (
10
) that is innitely sharp and innitely
narrow, but with unit area, to a real molecular lineshape. As described below, the
lineshape function g(
10
) is typically either a Lorentzian or a Gaussian function
with unit area but nite width and height, and now equations (1.52) and (1.53) are
replaced by
(B
10
)

=
2
2
3
0
h
2

2
10
g(
10
) (1.54)
and
(A
10
)

=
16
3

3
3
0
hc
3

2
10
g(
10
), (1.55)
respectively, in which
_
(B
10
)

d = B
10
and
_
(A
10
)d = A
10
. In practice the
subscripts in equations (1.54) and (1.55) are suppressed and the same symbols A
10
and B
10
are used with and without lineshape functions, although the dimensions are
dierent. In particular, as A

and A are related by A

= Ag(
10
), A has dimensions
s
1
and g(
10
) has dimensions s, so that A

is dimensionless. Note also that the two


levels 1 and 0 are assumed to be nondegenerate. The usual cases of degenerate atomic
and molecular energy levels will be considered in later chapters.
Beers Law
Consider a system (Figure 1.12) with N
0
molecules per cubic meter in the ground state
and N
1
in the excited state. A ux of photons F
0
= I
0
/h (units of photons m
2
s
1
)
is incident upon the system from the left. As these photons travel through the system
they can be absorbed or they can induce stimulated emission. What is the intensity of
radiation after a distance l ?
If only absorption and stimulated emission are considered, then at a particular
distance x we can write
20 1. Introduction
1m
1m
E
1
E
0
N
1
N
0
dx
l
I =Fh
0
n
0
Figure 1.12: A system with dimensions 1 m 1 m l m that contains molecules.
dN
1
dt
= B
10
N
1
+ B
10
N
0
=
2
2

2
10
3
0
h
2
(N
0
N
1
)g(
10
)
=
2
2

2
10

3
0
hc
(N
0
N
1
)g(
10
)F
= F(N
0
N
1
), (1.56)
in which = I/c = hF/c has been used. The absorption cross section is dened in
this way as
=
2
2

2
10
3
0
hc
g(
10
), (1.57)
with dimensions of m
2
. The physical interpretation of is as the eective area that
a molecule presents to the stream of photons of ux F. Notice that (1.57) and (1.53)
can be combined to give the convenient equation
=
A
2
g(
10
)
8
=

2
g(
10
)
8
sp
, (1.58)
which relates the cross section to the radiative lifetime
sp
1/A
10
of a transition
for a two-level system. The subscript sp on refers to spontaneous emission.
Care is required when radiative lifetimes are used for real multilevel systems because
any given level n can emit spontaneously to all lower levels, so that

sp
=
1

A
nj
. (1.59)
In other words the lifetime is related to the rates of all radiative rates connecting
the upper state |n > to all lower energy states |j >, rather than just
sp
= 1/A
10
.
Any nonradiative processes add additional rate terms to the sum in equation (1.59).
The individual A and B coecients, however, still obey the equations developed for a
two-level system.
If a ux F is incident to the left of a small element of thickness dx (Figure 1.12)
with cross-sectional area of 1 m
2
, then the change in ux caused by passing through
the element is
1.3 Interaction of Radiation with Matter 21
dF = F(N
0
N
1
)dx. (1.60)
Upon integrating over the absorption path, this becomes
_
F
F
0
dF
F
= (N
0
N
1
)
_
l
0
dx
or
ln(
F
F
0
) = ln(
I
I
0
) = (N
0
N
1
)l. (1.61)
Expression (1.61) can also be rewritten in the form
I = I
0
e
(N
0
N
1
)l
, (1.62)
which is equivalent to the commonly encountered decadic version of Beers law,
I = I
0
10
cl
. (1.63)
It is common to report in cm
2
, N in molecules per cm
3
, and l in cm rather than
the corresponding SI units. The units used in Beers law (1.63) are customarily moles
per liter for c, cm for l, and liter mole
1
cm
1
for the molar absorption coecient, .
Sometimes the cross section and concentration are combined to dene an absorption
coecient = (N
0
N
1
) for a system, in which case we write
I = I
0
e
l
, (1.64)
rather than (1.62).
Lineshape Functions
A real spectrum of a molecule, such as that for gaseous CO
2
(Figure 1.13), contains
many absorption features called lines organized into a band associated with a par-
ticular mode of vibration. For the spectrum illustrated in Figure 1.13 the lines are
associated with the antisymmetric stretching mode,
3
, of CO
2
. At high resolution the
spectrum seems to consist of very narrow features, but if the scale is expanded the lines
are observed to have denite widths and characteristic shapes. What are the possible
lineshape functions g(
10
) and what physical processes are responsible for these
shapes?
Lineshape functions fall into one of two general categories: homogeneous and in-
homogeneous. A homogeneous lineshape occurs when all molecules in the system have
identical lineshape functions. For example, if an atomic or molecular absorber in the
gas phase is subject to a high pressure, then all molecules in the system are found
to have an identical pressure-broadened lineshape for a particular transition. Pressure
broadening of a transition is said, therefore, to be a homogeneous broadening.
In contrast, if a molecule is dissolved in a liquid, then the disorder inherent in the
structure of the liquid provides numerous dierent solvent environments for the solute.
Each solute molecule experiences a slightly dierent solvent environment and therefore
has a slightly dierent absorption spectrum. The observed absorption spectrum (Figure
1.14) is made up of all of the dierent spectra for the dierent molecular environments;
it is said to be inhomogeneously broadened.
22 1. Introduction
Figure 1.13: A typical molecular spectrum, the antisymmetric stretching mode of carbon
dioxide. The weak bending hot band (see Chapter 7) is also present.
Figure 1.14: An inhomogeneously broadened line made up of many homogeneously broadened
components.
The most important example of gas phase inhomogeneous broadening occurs be-
cause of the MaxwellBoltzmann distribution of molecular velocities and is called
Doppler broadening. The dierent molecular velocities give the incident radiation a
frequency shift of = (1 v/c)
0
in the molecular frame of reference. This results in
slightly dierent spectra for molecules moving at dierent velocities and results in an
inhomogeneous lineshape.
Natural Lifetime Broadening
Consider a two-level system with an intrinsic lifetime
sp
seconds for the level at energy
E
1
for the spontaneous emission of radiation (Figure 1.15). The wavefunction that
1.3 Interaction of Radiation with Matter 23
Figure 1.15: Spontaneous emission in a two-level system.
describes the state of the system in the absence of electromagnetic radiation is given by
(t) = a
0

0
(t) + a
1

1
(t)
= a
0

0
e
iE
0
t/
+ a
1

1
e
iE
1
t/
(1.65)
where a
0
and a
1
are simply constants. Should the system be excited into this superpo-
sition state (for example, by a pulse of electromagnetic radiation), the dipole moment
of the system in this state is given by the expectation value of the dipole moment
operator as
M(t) = ||
=
10
(a

0
a
1
e
i
10
t
+ a
0
a

1
e
i
10
t
)
= Re(2
10
a
0
a

1
e
i
10
t
), (1.66)
assuming that the space-xed dipole moments
0
||
0
and
1
||
1
both vanish
in states |1 and |0. (NB: Nonzero values for the dipole moment are still possible in
the molecular frame.) The dipole moment of the system oscillates at the Bohr angular
frequency
10
as
M(t) = 2a
0
a
1

10
cos(
10
t), (1.67)
if a
0
and a
1
are chosen to be real numbers. A system in such a superposition state has
a macroscopic oscillating dipole in the laboratory frame (Figure 1.16).
Now if the population in the excited state decreases slowly in time (relative to the
reciprocal of the Bohr frequency) due to spontaneous emission, then the amplitude of
the oscillation will also decrease. This corresponds to a slow decrease in a
1
(equation
(1.67)) at a rate of /2 where = 1/
sp
= A
10
. Thus the oscillating dipole moment
is now
M(t) = M
0
e
t/2
cos(
10
t) (1.68)
as shown in Figure 1.17.
24 1. Introduction
M
t
Figure 1.16: Oscillating dipole moment of a system in a superposition state.
Figure 1.17: Slowly damped oscillating dipole moment.
What frequencies are associated with a damped cosine wave? Clearly the undamped
wave oscillates innitely at exactly the Bohr frequency
10
. The distribution of frequen-
cies F() present in a waveform f(t) can be determined by taking a Fourier transform,
i.e., as
F() =
_

f(t)e
it
dt. (1.69)
Note also that an arbitrary waveform f(t) can be written as a sum (integral) over plane
waves e
it
, each with amplitude F(), as
f(t) =
1
2
_

F()e
it
d, (1.70)
which is referred to as the inverse Fourier transform. Thus F() measures the amount
of each frequency required to synthesize f(t) out of sine and cosine functions (e
it
= cos t +i sin t). Taking the Fourier transform of the time-dependent part of M(t)
gives
1.3 Interaction of Radiation with Matter 25
F() =
_

e
t/2
cos(
10
t)e
it
dt
=
1
2
_

0
e
t/2
(e
i(
10
)t
+ e
i(+
10
)t
)dt
=
1
2
_
1
/2 + i(
10
)
+
1
/2 + i( +
10
)
_
(1.71)
for the decay process beginning at t = 0. The nonresonant term, i.e., the term containing
+
10
, is dropped because
10
and
10
, so that it is negligible in comparison
with the resonant term containing
10
(cf. the rotating wave approximation). With
this (rather good) approximation, F() becomes
F()
1
2
1
/2 + i(
10
)
. (1.72)
In the semiclassical picture an oscillating dipole moment radiates power at a rate pro-
portional to |
10
|
2
(i.e., A
10
|
10
|
2
) and the lineshape function, given by
|F()|
2
=
1
4
1
(/2)
2
+ (
10
)
2
(1.73)
is an unnormalized Lorentzian. Normalization requires that
_

g(
10
)d =
_

g(
10
)d = 1, (1.74)
so that the nal normalized Lorentzian lineshape functions are
g(
10
) =
/(2)
(/2)
2
+ (
10
)
2
, (1.75)
and
g(
10
) =

(/2)
2
+ (2)
2
(
10
)
2
. (1.76)
Note that g(
10
) and g(
10
) are related by
2g(
10
) = g(
10
).
Without spontaneous emission the lineshape function would be (
10
) since the
innite cosine wave oscillates at a frequency of exactly
10
. The decaying cosine wave
caused by spontaneous emission gives a Lorentzian function of nite width (Figure
1.18). At the peak center ( =
10
) we have g(
10
) = 4/, and the function drops
to half this value when
(2)
2
(
1/2

10
)
2
= (

2
)
2
or (
1/2

0
) =

4
.
The full width at half maximum (FWHM), represented by
1/2
, is given as

1/2
=

2
=
1
2
sp
, (1.77)
since = 1/
sp
. The Lorentzian lineshape function (
0
=
10
) can thus be expressed as
26 1. Introduction
Figure 1.18: A normalized Lorentzian function.
g(
0
) =

1/2
/(2)
(
1/2
/2)
2
+ (
0
)
2
(1.78)
in terms of the full width at half maximum. Note that some authors use the half width
at half maximum as a parameter rather than the full width.
The important result
1/2
= 1/(2
sp
) agrees with the Heisenberg time-energy
uncertainty principle Et , or
E
h

sp

1
2
, (1.79a)
or

1
2
sp
. (1.79b)
The spontaneous lifetime of the excited state means that the atom or molecule cannot
be found at E
1
for more than
sp
on average. This provides a fundamental limit on
the linewidth arising from the transition between the two states (Figure 1.19). For-
mula (1.79b) has been checked experimentally, for example in the case of the sodium
3
2
P
3/2
3
2
S
1/2
transition (one of the famous sodium D-lines) at 5 890

A. The ex-
perimentally measured lifetime
sp
= 16 ns and the observed homogeneous linewidth

1/2
= 10 MHz are consistent with equation (1.79b). The uncertainty principle there-
fore requires that if an excited state exists for only
sp
seconds on average, then the
1.3 Interaction of Radiation with Matter 27
Figure 1.19: The spontaneous lifetime
sp
gives the transition E
1
E
0
a nite linewidth.
energy level E
1
cannot be measured relative to E
0
with an accuracy that is greater
than
1/2
Hz.
The expression
1/2
= 1/(2) has widespread use in chemical physics. For ex-
ample, if H
2
O is excited by vacuum ultraviolet light, it can dissociate very rapidly:
H
2
O
h
HO + H. (1.80)
If the H
2
O molecule exists in a given excited electronic state for only one vibrational pe-
riod ( = 3 600 cm
1
corresponding to an OH stretch), then according to the Heisenberg
uncertainty principle the lifetime will be given by = 9.3 10
15
s = 9.3 femtosec-
onds (fs). Thus the width (FWHM) of a line in the spectrum will be
1/2
= 1.710
13
Hz or
1/2
= 570 cm
1
. A measurement of the homogeneous width of a particular
spectral line can thus provide an estimate of the lifetime of the excited state.
Pressure Broadening
The derivation of the pressure-broadening lineshape is a dicult problem because it
depends on the intermolecular potentials between the colliding molecules. However, a
simplied model within the semiclassical picture gives some estimation of the eect.
Consider the two-level system discussed in the previous section with the wavefunc-
tion written as a superposition state. The dipole moment oscillates at the Bohr fre-
quency except during a collision. If the collision is suciently strong, then the phase of
the oscillating dipole moment is altered in a random manner by the encounter. Let the
average time between collisions be T
2
(Figure 1.20). The innite cosine wave is broken
by successive collisions into pieces of average length T
2
. The eect of collisions will
be to convert the innitely narrow lineshape associated with an innitely long cosine
wave into a lineshape function of nite width. The application of Fourier transform
arguments (using autocorrelation functions
6
) to decompose the broken waveform into
frequency components results in a Lorentzian lineshape with a width (FWHM) given by

1/2
=
1
T
2
. (1.81)
28 1. Introduction
M
t
T
2
Figure 1.20: The phase of an oscillating dipole moment randomly interrupted by collisions.
Since the average time between collisions is proportional to the reciprocal of the
pressure, p, it therefore follows that the FWHM will be proportional to the pressure, i.e.,

1/2
= bp, (1.82)
with b referred to as the pressure-broadening coecient. The quantitative calculation of
b without recourse to experiment poses a dicult theoretical problem. Experimentally,
typical values for b are about 10 MHz per Torr of the pressure-broadening gas.
In general not only are the lines broadened by increasing pressure but they are also
shifted in frequency. These shifts are generally small, often less than 1 MHz/Torr, but
they become important when very precise spectroscopic measurements are to be made.
Doppler Broadening
Doppler broadening results in an inhomogeneous lineshape function. If the transition
has an intrinsic homogeneous lineshape g
H
(

0
) centered at

0
, then the inhomoge-
neous distribution function g
I
(

0
), centered at
0
, is required to describe the total
lineshape function g(
0
) according to the expression
g(
0
) =
_

g
I
(

0
)g
H
(

0
)d

0
. (1.83)
The distribution function g
I
(

0
) gives the probability that a system has a resonance
frequency in the interval

0
to

0
+ d

0
, i.e.,
dp = g
I
(

0
)d

0
. (1.84)
The lineshape integral (1.83) is referred to mathematically as a convolution of the
two functions g
I
and g
H
, as can be made more apparent by making the substitution
x =

0
,
g(
0
) =
_

g
I
(x)g
H
((
0
) x)dx. (1.85)
Commonly the homogeneous lineshape function g
H
is Lorentzian, while the inhomoge-
neous function g
I
is a Gaussian: the convolution of these two functions is called a Voigt
lineshape function (Figure 1.21).
1.3 Interaction of Radiation with Matter 29
Figure 1.21: The Voigt lineshape is a convolution of an inhomogeneous Gaussian lineshape
function with a homogeneous Lorentzian lineshape function.
Figure 1.22: Interaction of a plane electromagnetic wave with a moving atom.
The Voigt lineshape function is a general form that can include purely homogeneous
or purely inhomogeneous lineshapes as limiting cases. If the width of the inhomogeneous
part is much greater than that of the homogeneous part, that is, if
I

H
, then
g
H
(

0
) (

0
) and
g(
0
) =
_

g
I
(

0
)(

0
)d

0
= g
I
(
0
). (1.86)
Conversely if
I

H
, then g
I
(

0
) (

0
) and g(
0
) = g
H
(
0
).
Consider an atom with velocity v interacting with a plane wave with a wave vector
k. If k is parallel to v, then the atom sees a Doppler shifted frequency

= (1 v/c)
depending upon whether the atom is moving in a direction that is the same as () or
opposite to (+) that of the electromagnetic radiation (Figure 1.22). In general, it is
only the component of v along k (i.e., v cos ) that matters, so that

=
_
1
v k
c|k|
_
, (1.87)
neglecting a small relativistic correction
7
(second-order Doppler eect).
The Doppler eect can be viewed in two equivalent ways. In the frame of the atom
it is the frequency of the electromagnetic wave which has been shifted, with the atom at
30 1. Introduction
rest at the origin of the atomic coordinate system. Alternatively, in the xed laboratory
frame the electromagnetic wave is unshifted at , but the atomic resonance frequency

0
(of the atom moving at velocity v) has been shifted to the new value of

0
=

0
1 v/c
. (1.88)
All that is required to obtain a lineshape function is the distribution of velocities.
The distribution of molecular velocity components along a given axis (such as k) in
a gaseous system is given by the MaxwellBoltzmann distribution function
p
v
dv =
_
m
2kT
_
1/2
e
mv
2
/(2kT)
dv (1.89)
for particles of mass m at a temperature T. Using equation (1.89) and dv = (c/
0
)d

0
(obtained by taking dierentials of equation (1.87)) gives the normalized inhomogeneous
lineshape function
g
D
(
0
) =
1

0
_
mc
2
2kT
_
1/2
e
mc
2
(
0
)
2
/(2kT
2
0
)
. (1.90)
The FWHM,
D
, is easily shown to be

D
= 2
0
_
2kT ln(2)
mc
2
, (1.91)
or

D
= 7.2 10
7

0
_
T
M
, (1.92)
in which T is in K, M in atomic mass units u,
0
in cm
1
, and
D
in cm
1
. The
Doppler lineshape function is thus given by
g
D
(
0
) =
2

D
_
ln(2)

e
4 ln 2((
0
)/
D
)
2
(1.93)
in terms of the Doppler FWHM.
The Gaussian function is the bell-shaped curve well known in statistics. The
Gaussian lineshape function is more sharply peaked around
0
than the corre-
sponding Lorentzian lineshape function (Figure 1.23). Notice the much more extensive
wings on the Lorentzian function in comparison with the Gaussian function.
For example, the Doppler width of the 3
2
P
3/2
3
2
S
1/2
transition of the Na atom
at 300 K is already
D
= 0.044 cm
1
= 1 317 MHz, and is much larger than the
natural linewidth of 10 MHz. In addition, if the Na atom is surrounded by Ar atoms
at 1 Torr total pressure, then pressure broadening contributes 27 MHz to the total
homogeneous linewidth of 37 MHz.
8
Visible and ultraviolet transitions of gas phase
atoms and molecules typically display Doppler broadening at low pressure because the
inhomogeneous linewidth greatly exceeds the total homogeneous linewidth.
1.3 Interaction of Radiation with Matter 31
Figure 1.23: Normalized Gaussian and Lorentzian lineshape functions.
Laser
Source
v
d
Molecular
Figure 1.24: A laser beam interacts with a molecular beam inside a vacuum chamber.
Transit-Time Broadening
Consider the experimental arrangement of Figure 1.24 in which a laser beam of width d
is crossed at right angles with a beam of molecules moving at a speed of v in a vacuum
chamber. The molecules can only interact with the radiation for a nite time, = d/v,
called the transit time. The time corresponds to the time required by a molecule
in the molecular beam to cross through the laser beam. If the laser is considered to
be perfectly monochromatic, with frequency , then a molecule experiences an electric
eld as shown in Figure 1.25, assuming a constant weak light intensity from one side
of the laser beam to the other. If an intense laser beam is used in these experiments,
Rabi oscillations (discussed earlier) are observed.
Suppose an innitely long (in time) oscillating electric eld with an innitely nar-
row frequency distribution has been chopped into a nite length with a nite frequency
width. The nite time allowed for the laser-molecule interaction has resulted in a broad-
ening of the transition. As far as the observed lineshape is concerned, it does not matter
whether the molecular resonance is broadened or the frequency distribution of the ap-
plied radiation is increased. The result is the same: a broader line.
The frequency distribution associated with the electric eld of Figure 1.25 is deter-
mined, as for lifetime broadening, by taking the Fourier transform. Thus we write
32 1. Introduction
Figure 1.25: A molecule experiences an electromagnetic wave of nite length .
F() =
_

E
0
cos(
0
t)e
it
dt
=
E
0
2
_
/2
/2
(e
i
0
t
+ e
i
0
t
)e
it
dt
=
E
0
2
_
e
i(
0
)t
i(
0
)
+
e
i(
0
+)t
i(
0
+ )
_/2
/2
E
0
sin((
0
)/2)
(
0
)
, (1.94)
in which the nonresonant term in
0
+ has, as usual, been discarded. Since the intensity
of the light is proportional to |E|
2
, the unnormalized lineshape is proportional to
|F()|
2
=
sin
2
((
0
)/2)
(
0
)
2
. (1.95)
The corresponding normalized lineshape function
g(
0
) =
2

sin
2
((
0
)/2)
(
0
)
2
, (1.96)
or
g(
0
) =
1

sin
2
(2(
0
)/2)
(
0
)
2
, (1.97)
is shown in Figure 1.26. The FWHM of this function is about
1/2
= 0.89/. For
an atom traveling at 500 m/s through a laser beam of 1 mm width, = 2 10
6
s
and
1/2
= 0.45 MHz. Although transit-time broadening is relatively small, it is not
negligible for very precise Doppler-free measurements or for microwave-molecular beam
measurements.
1.3 Interaction of Radiation with Matter 33
Figure 1.26: A plot of the lineshape function caused by transit-time broadening.
Power Broadening
The high-power laser has become an ubiquitous tool of modern spectroscopy. The ap-
plication of intense electromagnetic radiation to a system will cause the spectral lines
to broaden and even to split. The detailed calculation of a molecular lineshape at high
power is complicated, but a simple estimate of the linewidth is available from the time-
energy uncertainty principle of equation (1.48), i.e.,
Et .
At high powers the molecular system undergoes Rabi oscillations at an angular
frequency
R
=
10
E/. The system is thus in the excited state E
1
only for a period
of about h/
10
E, which therefore provides an estimate for t. Using this estimate for
t in the uncertainty relation (1.48) gives
E
h

10
E

or equivalently, using h = E, a frequency uncertainty of


10
E
2h
=

R
4
2
. (1.98)
For example, a 1-W laser beam of 1-mm diameter interacting with a two-level system
with a transition dipole moment of 1 D and a Rabi frequency
R
= 9.8 10
8
rad/s
already gives = 25 MHz. Pulsed lasers can easily achieve peak powers of 1 MW
(10 mJ in 10 ns), which will increase E, equation (1.43), by a factor of 1 000, to 3.1
10
7
V/m for the preceding example. The power-broadened linewidth is then of the order
25 000 MHz or 0.83 cm
1
, which is larger than a typical Doppler width for a visible or
UV electronic transition.
34 1. Introduction
In this chapter the interaction of light with matter has been discussed and key
equations derived. For spectroscopy perhaps the most important equations are (1.57),
(1.58), and (1.62) because they relate the microscopic molecular world to macroscopic
absorption and emission rates. The transition dipole moment
M
10
=
10
=
_

0
d (1.99)
will be used numerous times to derive selection rules. The absorption strength, equation
(1.54), of a transition and the emission rate, equation (1.55), are proportional to the
square of the transition dipole moment. The absorption (I/I
0
) associated with a par-
ticular molecular transition depends upon three factors: the absorption cross section ,
the population dierence between the two levels, N
0
N
1
, and the optical path length
l in equation (1.62), namely,
I
I
0
= e
(N
0
N
1
)l
.
Spectroscopists use these relationships constantly.
Problems
1. The helium-neon laser operates at wavelength 6 328.165

A in air. The refractive
index of air is 1.000 275 9 at this wavelength.
(a) What is the speed of light in air and the vacuum wavelength?
(b) What is the wavenumber (cm
1
) and the magnitude of the wavevector in
air.
(c) What are the frequency and the period of oscillation?
(d) Calculate the energy and momentum of a photon with an air wavelength of
6 328.165

A.
(e) What will be the wavelength and frequency in water if the refractive index
is 1.333 3?
2. The refractive index of dry air at 15

C and 760 Torr pressure is given by the


Cauchy formula
10
7
(n 1) = 2 726.43 + 12.288
10
8

2
+ 0.355 5
10
16

4
where is in

A (CRC Handbook of Chemistry and Physics, CRC Press). (The
formula due to Edlen, Metrologia 2, 71 (1966) is slightly more accurate but less
convenient to use.)
(a) Calculate the refractive index of air at 4 000

A, 6 000

A, and 8 000

A.
(b) Convert the air wavelengths into vacuum wavelengths and calculate the cor-
responding wavenumbers (cm
1
).
3. (a) What is the momentum and the de Broglie wavelength associated with a
human weighing 150 lb and walking at 4 miles/hr?
Problems 35
(b) What is the momentum and the de Broglie wavelength of an electron accel-
erated through a voltage of 100 V?
4. A crystal lattice has a typical spacing of 2

A.
(a) What velocity, momentum, and kinetic energy should be used for the elec-
trons in an electron diraction experiment? (Hint: d (lattice spacing)
for a good diraction experiment.)
(b) What voltage needs to be applied to the electron gun for the diraction
experiment?
(c) Answer part (a) if neutrons were used instead of electrons.
(d) If the diraction experiment were carried out with photons, then what wave-
length, energy, and momentum would be required?
5. (a) Make the necessary conversions in order to ll in the table:
Wavelength (

A) 420
Wavenumber (cm
1
) 100
Energy (J)
Energy (kJ/mole) 490
Frequency (Hz) 8.21 10
13
(b) Name the spectral region associated with each of the last four columns of
the table.
6. There are two limiting cases associated with the Planck function

(T) =
8h
3
c
3
1
e
h/kT
1
. (1.100)
(a) Calculate kT at room temperature (20

C) in J, kJ/mole, eV, and cm


1
.
(b) For long wavelengths (microwave frequencies at room temperature) h
kT. In this case, derive a simpler, approximate equation (called the Rayleigh
Jeans law) for

(T). This is the formula derived using classical arguments


prior to Plancks quantized oscillator approach.
(c) For high-energy photons (near infrared wavelengths at room temperature)
h kT. In this case, derive a simpler, approximate expression for

(T)
called Wiens formula.
7. (a) Dierentiate the Planck function to determine the frequency at which

is
a maximum (Figure 1.5).
(b) Convert the Planck law from a function of frequency to a function of wave-
length; that is, derive

d from

d.
(c) Derive the Wien displacement law for blackbody radiation

max
T = 2.898 10
3
mK (1.101)
using

d.
36 1. Introduction
(d) What wavelengths correspond to the maximum of the Planck function in
interstellar space at 3 K, at room temperature (20

C), in a ame (2 000

C),
and in the photosphere of the sun (6 000 K)?
8. The total power at all frequencies emitted from a small hole in the wall of a
blackbody cavity is given by the StefanBoltzmann law
I = T
4
,
where = 5.670 5 10
8
W m
2
K
4
is the StefanBoltzmann constant.
(a) Derive the StefanBoltzmann law.
(b) Determine an expression for in terms of fundamental physical constants
and obtain a numerical value. (Hint:
_

0
x
3
/(e
x
1)dx =
4
/15.)
9. Derive the Wien displacement law (Problem 7) using

rather than

.
10. Consider the two-level system (Figure 1.6) at room temperature, 20

C, and in
the photosphere of the sun at 6 000 K. What are the relative populations N
1
/N
0
corresponding to transitions that would occur at 6 000

A, 1 000 cm
1
, 100 GHz,
and 1 GHz?
11. A 100-W tungsten lament lamp operates at 2 000 K. Assuming that the lament
emits as a blackbody, what is the total power emitted between 6 000

A and
6 001

A? How many photons per second are emitted in this wavelength interval?
12. (a) What is the magnitude of the electric eld for the beam of a 1-mW helium-
neon laser with a diameter of 1 mm?
(b) How many photons per second are emitted at 6 328

A?
(c) If the laser linewidth is 1 kHz, what temperature would a blackbody have to
be at in order to emit the same number of photons from an equal area over
the same frequency interval as the laser?
13. Derive the relationship (1.13) between the energy density

and the intensity I

for a blackbody
I

c
o
/4.
(Hint: The total power passing through the hole and present in the solid angle
d is

c cos (d/4).)
14. Solve equations (1.36a) and (1.36b) for the interaction of light with a two-level
system.
(a) First convert the two rst-order simultaneous dierential equations into a
single second-order equation by substituting one equation (1.36a) into the
other (1.36b).
Problems 37
(b) The general solution for a second-order dierential equation with constant
coecients is
a
0
(t) = Ae
it
+ Be
it
.
Show that this form for a
0
(t) implies that a
1
(t) is given by
a
1
(t) =
2
i
R
e
it
(Aie
it
+ Bie
it
).
(c) To obtain and substitute the general solution into the second-order
equation and obtain the characteristic equation

2


2
R
4
= 0.
The two solutions of this equation are and . Find and , and simplify
your answer using the denition
(
2
R
+
2
)
1/2
.
(d) Determine A and B from the initial conditions at t = 0, a
0
(0) = 1, and
a
1
(0) = 0, and derive equations (1.37) and (1.38).
(e) Verify that the nal answer (1.37) and (1.38) satises the dierential equa-
tions (1.36a) and (1.36b).
15. The 3
2
P
3/2
3
2
S
1/2
transition of Na (actually the (F = 3) (F = 2) hyperne
transition) has a Rabi frequency of 4.15 10
8
rad/s with a laser intensity of 560
mW/cm
2
. What is the transition dipole moment in debye?
16. The lifetime of the 3
2
P
1/2
3
2
S
1/2
transition of the Na atom at 5 896

A is
measured to be 16.4 ns.
(a) What are the Einstein A and B coecients for the transition?
(b) What is the transition dipole moment in debye?
(c) What is the peak absorption cross section for the transition in

A
2
, assuming
that the linewidth is determined by lifetime broadening?
17. What are the Doppler linewidths (in cm
1
) for the pure rotational transition of
CO at 115 GHz, the infrared transition of CO
2
at 667 cm
1
, and the ultraviolet
transition of the Hg atom at 2 537

A, all at room temperature (20

C)?
18. Calculate the transit-time broadening for hydrogen atoms traversing a 1-mm
diameter laser beam. For the speed of the hydrogen atoms use the rms speed
(v = (3kT/m)
1/2
) at room temperature (20

C).
19. At what pressure will the Doppler broadening (FWHM) equal the pressure broad-
ening (FWHM) for a room temperature (20

C) sample of CO gas for a pure rota-


tional transition at 115 GHz, a vibration-rotation transition at 2 140 cm
1
, and an
electronic transition at 1 537

A? Use a typical pressure-broadening coecient
of 10 MHz/Torr in all three cases.
38 1. Introduction
20. What are the minimum spectral linewidths (in cm
1
) of pulsed lasers with pulse
durations of 10 fs, 1 ps, 10 ns, and 1 s?
21. (a) For Na atoms in a ame at 2 000 K and 760-Torr pressure, calculate the peak
absorption cross section (at line center) for the 3
2
P
1/2
3
2
S
1/2
transition
at 5 896

A. Use 30 MHz/Torr as the pressure-broadening coecient and the
data in Problem 16.
(b) If the path length in the ame is 10 cm, what concentration of Na atoms
will produce an absorption (I/I
0
) of 1/e at line center?
(c) Is the transition primarily Doppler or pressure broadened?
(d) Convert the peak absorption cross section in cm
2
to the decadic peak molar
absorption coecient (see equation (1.63)).
22. For Ar atoms at room temperature (20

C) and 1-Torr pressure, estimate a colli-


sion frequency for an atom from the van der Waals radius of 1.5

A. What is the
corresponding pressure-broadening coecient in MHz/Torr?
23. A stationary atom of mass m emits a photon of energy h and momentum k.
(a) Use the laws of conservation of energy and momentum to show that the shift
in frequency of the emitted photon due to recoil of the atom is given by

0

h
0
2mc
2
.
(b) What is the shift in frequency due to recoil of the atom for the Na D line at
5 890

A?
(c) What is the shift in frequency for a -ray of energy 1 369 keV emitted from
24
Na?
24. At the top of the earths atmosphere the solar irradiance is 1 368 W/m
2
(the solar
constant). Calculate the magnitude of the electric eld, assuming a plane wave
at a single frequency for E.
25. At night, the concentration of the NO
3
free radical is about 10
9
molecules/cm
3
near the ground. NO
3
has a visible absorption band near 662 nm, with a peak
absorption cross section of 2.3 10
17
cm
2
molecule
1
at 298 K. For an absorp-
tion path of 1 km, what will be the change in atmospheric transmission (1I/I
0
)
at 662 nm due to NO
3
?
26. For transit time broadening, consider the typical case (Figure 1.24) of a molecular
beam crossing through a Gaussian laser beam, i.e., the applied electric eld is
given as
E = E
0
e
(t/a)
2
cos
0
t.
(a) What is the normalized lineshape function, g(
0
)?
(Hint:
_

e
p
2
x
2
qx
dx =

p
e
q
2
/4p
2
, p > 0.)
(b) What is the full width at half maximum
1/2
of g(
0
)?
References 39
(c) What is the t product for a Gaussian beam? How does this value com-
pare with that of the corresponding Heisenberg uncertainty principle?
(d) The radial distribution of the electric eld of a Gaussian laser beam is pro-
portional to e
(r/w)
2
. What is the relationship between the Gaussian beam
width parameter w and the parameter a dened above?
References
1. Mills, I., Cvitas, T., Homann, K., Kallay, N., and Kuchitsu, K., Quantities, Units
and Symbols in Physical Chemistry, 2nd ed., Blackwell, Oxford, 1993; see also
http://www.iupac.org/.
2. Mohr, P. J. and Taylor, B. N., CODATA Recommended Values of the Fun-
damental Physical Constants: 1998, Rev. Mod. Phys. 72, 351 (2000); see
http://www.codata.org/.
3. Audi, G., Wapstra, A. H., and Thibault, C., Nucl. Phys. A. 729, 337 (2003); see
http://ie.lbl.gov/.
4. Lide, D. R., editor, CRC Handbook of Chemistry and Physics, 85th ed., CRC
Press, Boca Raton, Florida, 2004.
5. Bass, M., Van Stryland, E. W., Williams, D. R., and Wolfe, W. L., editors, Hand-
book of Optics, Vol. II, McGrawHill, New York, 1995, Chapter 24.
6. Svelto, O., Principles of Lasers, 4th ed., Plenum, New York, 1998, pp. 44-46,
Appendix B.
7. Demtroder, W., Laser Spectroscopy, 3rd ed., SpringerVerlag, Berlin, 2002, p. 768.
8. Demtroder, W., Laser Spectroscopy, 3rd ed., SpringerVerlag, Berlin, 2002, p. 77.
General References
Andrews, D. L. and Demidov, A. A., An Introduction to Laser Spectroscopy, 2nd
ed., Kluwer, New York, 2002.
Bracewell, R. N., The Fourier Transform and Its Applications, 3rd ed., McGraw
Hill, New York, 1999.
Corney, A., Atomic and Laser Spectroscopy, Oxford University Press, Oxford,
1977.
Demtroder, W., Laser Spectroscopy, 3rd ed., SpringerVerlag, Berlin, 2002.
Fowles, G. R., Introduction to Modern Optics, 2nd ed., Dover, New York, 1989.
Letokhov, V. S. and Chebotayev, V. P., Nonlinear Laser Spectroscopy, Springer
Verlag, Berlin, 1977.
Levenson, M. D. and Kano, S. S., Introduction to Nonlinear Laser Spectroscopy,
2nd ed., Academic Press, San Diego, 1988.
40 1. Introduction
Milonni, P. W. and Eberly, J. H., Lasers, Wiley, New York, 1988.
Siegman, A. E., Lasers, University Science Books, Mill Valley, California, 1986.
Steinfeld, J. I., Molecules and Radiation, 2nd ed., MIT Press, Cambridge, 1985.
Svelto, O., Principles of Lasers, 4th ed., Plenum, New York, 1998.
Yariv, A., Quantum Electronics, 3rd ed., Wiley, New York, 1989.

You might also like